Enantioselective Synthesis of Functionalized 4-Aryl ... - ACS Publications

Jun 5, 2017 - Acrylamide in Rauhut-Currier reaction; intramolecular isomerization of ... Phosphine-catalyzed intramolecular Rauhut–Currier reaction:...
1 downloads 0 Views 972KB Size
Letter pubs.acs.org/OrgLett

Enantioselective Synthesis of Functionalized 4‑Aryl Hydrocoumarins and 4‑Aryl Hydroquinolin-2-ones via Intramolecular Vinylogous Rauhut−Currier Reaction of para-Quinone Methides Xiang-Zhi Zhang,† Kang-Ji Gan,† Xiao-Xue Liu,† Yu-Hua Deng,† Fang-Xin Wang,† Ke-Yin Yu,† Jing Zhang,† and Chun-An Fan*,†,‡ †

State Key Laboratory of Applied Organic Chemistry, College of Chemistry and Chemical Engineering, Lanzhou University, 222 Tianshui Nanlu, Lanzhou 730000, China ‡ Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin 300071, China S Supporting Information *

ABSTRACT: A novel strategy for the asymmetric construction of functionalized 4-aryl-3,4-dihydrocoumarins and 4-aryl-3,4-dihydroquinolin-2-ones via an intramolecular vinylogous Rauhut−Currier reaction of para-quinone methides (p-QMs) under the bifunctional catalysis of chiral amine-phosphine is described. This intramolecular mode for the catalytic enantioselective 1,6-conjugate addition of pQMs has been explored for the first time, delivering two types of synthetically important heterocycles in high yields and enantioselectivites.

T

Scheme 1. Catalytic Enantioselective Approaches

he heterocyclic skeleton with 4-aryl-3,4-dihydrocoumarin or 4-aryl-3,4-dihydroquinolin-2-one exists in many biologically active natural products and pharmaceuticals,1 exhibiting anticancer1c and antibacterial1e activities (Figure 1). These

Figure 1. Selected bioactive natural products and pharmaceuticals.

structurally unique scaffolds are also common building blocks in natural product synthesis.2 Owing to their promising biological profiles as well as synthetic utilities, numerous protocols for the assembly of 4-aryl-3,4-dihydrocoumarins and 4-aryl-3,4-dihdroquinolin-2-ones have been reported.3,4 Among them, only a handful of approaches have been developed for the catalytic asymmetric version. For example, rhodium-catalyzed 1,4addition of arylboronic acids with coumarins5 or asymmetric hydrogenation of 4-aryl coumarins6 has provided straightforward routes to such types of 4-aryl-3,4-dihydrocoumarins (Scheme 1a). Recently, the Hou group has realized the kinetic resolution of 4-substituted-3,4-dihydrocoumarins via a palladium-catalyzed allylic substitution for optically active trans-3,4-disubstituteddihydrocoumarin derivatives7 (Scheme 1a). In addition, these skeletons could be also accessed by [4 + 2] annulation of oquinone methides with different two-carbon synthons in the presence of chiral organocatalysts8 (Scheme 1b). Meanwhile, Nheterocyclic carbene-catalyzed annulation of phenols with enals or alkynals has been demonstrated as a valuable method for the generation of 4-substituted-3,4-dihydrocoumarins9 (Scheme © 2017 American Chemical Society

1b). To our knowledge, however, only two reports have achieved the catalytic asymmetric synthesis of 4-aryl-3,4-dihydroquinolin2-ones through rhodium/palladium-catalyzed reaction of arylboronic acids with acrylamides10a and palladium-catalyzed cyclocarbonylation of 2-vinylanilines,10b respectively (Scheme 1c). Thus, it is highly desirable to explore alternative catalytic enantioselective methodologies to construct both 4-aryl-3,4dihydrocoumarins and 4-aryl-3,4-dihydroquinolin-2-ones. para-Quinone methides (p-QMs)11,12 as vinylogous Michael acceptors have attracted much attention in the synthetic community. To date, 1,6-conjugate addition,13,14 [2 + 1]annulation,15 [3 + 2]-annulation,16 [4 + 1]-annulation,17 and [4 Received: May 3, 2017 Published: June 5, 2017 3207

DOI: 10.1021/acs.orglett.7b01331 Org. Lett. 2017, 19, 3207−3210

Letter

Organic Letters + 2]-annulation 18 constitute the main context of the methodological studies on p-QMs. With our continuing interest in the chemistry of p-QMs,14a,j,k,15c,17a the installation of an electrophilic allyl ester or allyl amide group at the ortho position on the aromatic ring of p-QMs might render the possibility of a new intramolecular Rauhut−Currier reaction,19,20 which would lead to access to the synthetically important, optically active heterocycles (Scheme 1d). Recently, two reports involving intermolecular Rauhut−Currier-type 1,6-conjugate addition of p-QMs with vinyl ketones have been presented by Liu and Zhang21a and Wu.21b Herein, we described an enantioselective synthesis of functionalized 4-aryl-3,4-dihydrocoumarins and 4aryl-3,4-dihydroquinolin-2-ones through an unprecedented intramolecular vinylogous Rauhut−Currier reaction of p-QMs under the catalysis of nucleophilic chiral phosphines. We initially conducted the model reaction using p-QM ester 1a in the presence of various chiral phosphine catalysts 3a−3i (Table 1). The C2-symmetric chiral catalysts (R)-BINAP 3a and

With the optimized conditions in hand, we then explored the substrate scope (Scheme 2). A series of p-QM esters 1a−1m with Scheme 2. Substrate Scope of p-QM Esters

Table 1. Optimization of Reaction Conditionsa

entry

catalyst

time (h)

yield (%)b

ee (%)c

1 2 3 4 5 6 7 8 9

3a ((R)-BINAP) 3b ((R,R)-DIOP) 3c 3d 3e 3f 3g 3h 3i

24 12 24 24 24 10 2.5 1 24

NR 71 85 80 NR 85 98 98 NR

ND −24 24 24 ND 72 92 98 ND

electronically different substituents on aromatic ring A were first examined, giving the corresponding 4-aryl hydrocoumarins 2a− 2m in high yields (89−99% yield) with moderate-to-excellent enantioselectivities (50−99% ee). Notably, while using p-QM esters 1b and 1n with sterically bulky ortho-substituents, the reaction proceeded smoothly to afford 2b (99% yield) and 2n (99% yield), respectively, but with a dramatic decrease of enantioselectivities (low to 13% ee). The absolute configuration of 2e was unambiguously assigned by X-ray crystallographic analysis.23 In addition, p-QM ester 1o with a less bulky isopropyl group at the α,α′-position could afford desired product 2o in high yield and enantioselectivity (92% yield, 97% ee). Moreover, a stereoisomeric mixture of 1p (R1 = Me, R2 = t-Bu, E/Z = 1.5:1) was also examined, and product 2p was readily obtained without a negative influence on both reactivity and enantioselectivity (95% yield, 94% ee). To further expand the substrate scope of this methodology, various p-QM amides 4a−4e were then probed (Scheme 3). Compared with cases using p-QM esters (Scheme 2), the reaction of p-QM amides with the electron withdrawing N-Ts group could generally be completed in 15 min, showing a significant increase in reactivity. p-QM amides containing electron-neutral (R = H), electron-withdrawing (R = F), or electron-donating substituents (R = Me, OMe) at the C6 position of aromatic ring A could afford the desired 4-aryl-3,4dihydroquinolin-2-ones 5a−5d in high yields (up to 98% yield) and excellent enantioselectivities (up to 94% ee). The absolute stereochemsitry of 5a was clearly determined by X-ray crystallographic analysis.23 Moreover, when p-QM amide 4e bearing a C7−CF3 group was exposed to the optimized conditions, product 5e could be furnished with a satisfactory result (91% yield, 94% ee). It should be noted that substrate 4f with an electron-donating N-methyl group did not undergo the expected intramolecular cyclization, mostly due to the highly decreased electrophilicity of the corresponding N-methylated acrylamide moiety.

a Performed with 1a (0.1 mmol) and catalyst 3 (0.01 mmol) in CH2Cl2 (2.0 mL) at 15 °C. bYield of isolated products. cDetermined by HPLC.

(R,R)-DIOP 3b were first tested in this reaction (entries 1 and 2). Despite no reactivity observed using atropisomeric chiral bis(triaryl)phosphine 3a, the desired hydrocoumarin 2a could be obtained in 71% yield, albeit with 24% ee, under the catalysis of centrally chiral phosphine 3b. To improve the enantioselectivity, we then examined a series of bifunctional aminephosphine catalysts 3c−3i (entries 3−9). As compared with negative results in the cases employing 1,3-amine-phosphines 3c−3e (entries 3−5), amino acid-derived 1,2-amine-phosphine catalyst 3f gave dramatically improved enantiocontrol with 72% ee (entry 6). Encouraged by this promising result, two N-Boc1,2-amine-phosphines 3g and 3h22 were further investigated (entries 7 and 8). To our delight, catalyst 3h with increasing size of the substituent on the chiral center delivered the optimal result (98% yield, 98% ee, entry 8). Notably, cyclic trans-1,2-aminephosphine catalyst 3i was also surveyed (entry 9), but surprisingly, there was no reaction observed in this case. 3208

DOI: 10.1021/acs.orglett.7b01331 Org. Lett. 2017, 19, 3207−3210

Letter

Organic Letters Scheme 3. Substrate Scope of p-QM Amides

in the presence of an increased amount of AlCl3 (15 equiv) at an enhanced temperature (50 °C), the formation of unusual rearrangement product 6e could be improved to 72% yield, to some extent demonstrating the positive influence of reaction temperature on such [3.3] sigmatropic rearrangement. On the basis of the crystallographically determined absolute configurations of products 2e and 5a, a tentative model was proposed for the enantiocontrol observed in this reaction (Figure 2). Mechanistically, initial nucleophilic attack of 1,2-amine-

Figure 2. Proposed model for the enantioselectivity.

phosphine catalyst 3h on the acrylate moiety of p-QMs provided the zwitterionic intermediates A stabilized by hydrogen bonding interaction, and subsequently, a stereoselective 1,6-addition of less hindered Re face of enolate motif to Si face of p-QM unit could predominantly occur to give desired chiral products 2 and 5. In conclusion, a highly catalytic enantioselective method for the effective synthesis of chiral-functionalized coumarins and quinolinones has been developed via an intramolecular vinylogous Rauhut−Currier reaction of p-QMs under the bifunctional catalysis of nucleophilic chiral amine-phosphines. A series of synthetically useful, optically active 4-aryl-3,4-dihydrocoumarins and 4-aryl-3,4-dihydroquinolin-2-ones were achieved in high yields (up to 99% yield) and excellent enantioselectivities (up to 99% ee). This is the first example of a catalytic enantioselective 1,6-addition reaction of p-QMs in an intramolecular fashion, the utility of which has been manifested by the practicability of scaleup and the diverse transformations of heterocyclic products. The current exploration of introducing additional electrophilic sites provides a new perspective for methodological design concerning the chemistry of p-QMs.

For the synthetic utility of this protocol to be demonstrated, a gram-scale reaction of 1a was performed with a lower catalyst loading of 3h (0.05 equiv) (Scheme 4, route a), and the Scheme 4. Gram-Scale Synthesis and Its Transformations



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.orglett.7b01331. Experimental procedures and spectral data (PDF) X-ray data for compounds 2a (CIF), 5a (CIF), and 6e (CIF)

hydrocoumarin 2a was readily obtained (98% yield, 98% ee). With 2a in hand, its transformations were then explored. A facially selective hydrogenation of 2a (route b) led to the formation of desired dihydrocoumarin 6a with vicinal stereocenters (90% yield, >20:1 dr, 98% ee),24 and pleasingly, AlCl3mediated de-tert-butylation of 6a (route c) could smoothly give 7a (92% yield, 98% ee). Moreover, a diastereoselective cyclopropanation of 2a with 2-bromomalonic ester (route d) proceeded to deliver the spirocycle 6b (99% yield, >20:1 dr, 98% ee).24 Furthermore, when 2a was subjected to NaOMepromoted alcoholysis (route e), the resulting acyclic product 6c could be obtained in 99% yield without detectable racemization (97% ee) at the diarylmethine stereocenter. Upon treating 2a with 4 equiv of AlCl3 at 15 °C (route f), the expected de-tert-butylation product 6d was formed in 44% yield and 98% ee, but interestingly, an unexpected product 6e resulting from a [3.3] sigmatropic rearrangement was isolated in 50% yield. The structure of 6e was further assigned by X-ray crystallographic analysis.23 Notably, while this de-tert-butylation was conducted



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Chun-An Fan: 0000-0003-4837-3394 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful for financial support from NSFC (21572083, 21322201, 21290180), FRFCU (lzujbky-2015-48, lzujbky-2016ct02, lzujbky-2016-ct07), PCSIRT (IRT_15R28), the 111 Project of MOE (111-2-17), and the Chang Jiang Scholars 3209

DOI: 10.1021/acs.orglett.7b01331 Org. Lett. 2017, 19, 3207−3210

Letter

Organic Letters

13711. (e) Dong, N.; Zhang, Z.-P.; Xue, X.-S.; Li, X.; Cheng, J.-P. Angew. Chem., Int. Ed. 2016, 55, 1460. (f) Jarava-Barrera, C.; Parra, A.; López, A.; Cruz-Acosta, F.; Collado-Sanz, D.; Cárdenas, D. J.; Tortosa, M. ACS Catal. 2016, 6, 442. (g) Zhao, K.; Zhi, Y.; Wang, A.; Enders, D. ACS Catal. 2016, 6, 657. (h) He, F.-S.; Jin, J.-H.; Yang, Z.-T.; Yu, X.; Fossey, J. S.; Deng, W.-P. ACS Catal. 2016, 6, 652. (i) Li, X.; Xu, X.; Wei, W.; Lin, A.; Yao, H. Org. Lett. 2016, 18, 428. (j) Deng, Y.-H.; Zhang, X.-Z.; Yu, K.Y.; Yan, X.; Du, J.-Y.; Huang, H.; Fan, C.-A. Chem. Commun. 2016, 52, 4183. (k) Zhang, X.-Z.; Deng, Y.-H.; Yan, X.; Yu, K.-Y.; Wang, F.-X.; Ma, X.-Y.; Fan, C.-A. J. Org. Chem. 2016, 81, 5655. (l) Wong, Y. F.; Wang, Z.; Sun, J. Org. Biomol. Chem. 2016, 14, 5751. (m) Ge, L.; Lu, X.; Cheng, C.; Chen, J.; Cao, W.; Wu, X.; Zhao, G. J. Org. Chem. 2016, 81, 9315. (n) Chen, M.; Sun, J. Angew. Chem., Int. Ed. 2017, 56, 4583. (15) For [2 + 1]-annulations of p-QMs, see: (a) Yuan, Z.; Fang, X.; Li, X.; Wu, J.; Yao, H.; Lin, A. J. Org. Chem. 2015, 80, 11123. (b) Gai, K.; Fang, X.; Li, X.; Xu, J.; Wu, X.; Lin, A.; Yao, H. Chem. Commun. 2015, 51, 15831. (c) Zhang, X.-Z.; Du, J.-Y.; Deng, Y.-H.; Chu, W.-D.; Yan, X.; Yu, K.-Y.; Fan, C.-A. J. Org. Chem. 2016, 81, 2598. (d) Roiser, L.; Waser, M. Org. Lett. 2017, 19, 2338. (16) For recent selected examples on [3 + 2]-annulations of p-QMs, see: (a) Yuan, Z.; Wei, W.; Lin, A.; Yao, H. Org. Lett. 2016, 18, 3370. (b) Ma, C.; Huang, Y.; Zhao, Y. ACS Catal. 2016, 6, 6408. (17) For [4 + 1]-annulations of p-VQMs, see: (a) Zhang, X.-Z.; Deng, Y.-H.; Gan, K.-J.; Yan, X.; Yu, K.-Y.; Wang, F.-X.; Fan, C.-A. Org. Lett. 2017, 19, 1752. (b) Yuan, Z.; Gai, K.; Wu, Y.; Wu, J.; Lin, A.; Yao, H. Chem. Commun. 2017, 53, 3485. (18) For recent selected example on [4 + 2]-annulations of p-QMs, see: Zhao, K.; Zhi, Y.; Shu, T.; Valkonen, A.; Rissanen, K.; Enders, D. Angew. Chem., Int. Ed. 2016, 55, 12104. (19) For reviews, see: (a) Methot, J. L.; Roush, W. R. Adv. Synth. Catal. 2004, 346, 1035. (b) Aroyan, C. E.; Dermenci, A.; Miller, S. J. Tetrahedron 2009, 65, 4069. (c) Basavaiah, D.; Reddy, B. S.; Badsara, S. S. Chem. Rev. 2010, 110, 5447. (d) Wei, Y.; Shi, M. Chem. Rev. 2013, 113, 6659. (e) Xie, P.; Huang, Y. Eur. J. Org. Chem. 2013, 2013, 6213. (f) Bharadwaj, K. C. RSC Adv. 2015, 5, 75923. (20) For selected examples of asymmetric intramolecular Rauhut− Currier reactions, see: (a) Aroyan, C. E.; Miller, S. J. J. Am. Chem. Soc. 2007, 129, 256. (b) Marqués-López, E.; Herrera, R. P.; Marks, T.; Jacobs, W. C.; Könning, D.; de Figueiredo, R. M.; Christmann, M. Org. Lett. 2009, 11, 4116. (c) Aroyan, C. E.; Dermenci, A.; Miller, S. J. J. Org. Chem. 2010, 75, 5784. (d) Gong, J.-J.; Li, T.-Z.; Pan, K.; Wu, X.-Y. Chem. Commun. 2011, 47, 1491. (e) Selig, P. S.; Miller, S. J. Tetrahedron Lett. 2011, 52, 2148. (f) Wang, X.-F.; Peng, L.; An, J.; Li, C.; Yang, Q.-Q.; Lu, L.-Q.; Gu, F.-L.; Xiao, W.-J. Chem. - Eur. J. 2011, 17, 6484. (g) Takizawa, S.; Nguyen, T. M.-N.; Grossmann, A.; Enders, D.; Sasai, H. Angew. Chem., Int. Ed. 2012, 51, 5423. (h) Zhang, X.-N.; Shi, M. Eur. J. Org. Chem. 2012, 2012, 6271. (i) Osuna, S.; Dermenci, A.; Miller, S. J.; Houk, K. N. Chem. - Eur. J. 2013, 19, 14245. (j) Takizawa, S.; Nguyen, T. M.N.; Grossmann, A.; Suzuki, M.; Enders, D.; Sasai, H. Tetrahedron 2013, 69, 1202. (k) Su, X.; Zhou, W.; Li, Y.; Zhang, J. Angew. Chem., Int. Ed. 2015, 54, 6874. (l) Zhao, X.; Gong, J.-J.; Yuan, K.; Sha, F.; Wu, X.-Y. Tetrahedron Lett. 2015, 56, 2526. (m) Yao, W.; Dou, X.; Wen, S.; Wu, J.; Vittal, J. J.; Lu, Y. Nat. Commun. 2016, 7, 13024. (21) (a) Li, S.; Liu, Y.; Huang, B.; Zhou, T.; Tao, H.; Xiao, Y.; Liu, L.; Zhang, J. ACS Catal. 2017, 7, 2805. (b) Kang, T.-C.; Wu, L.-P.; Yu, Q.W.; Wu, X.-Y. Chem. - Eur. J. 2017, 23, 6509. (22) For the preparation of 3h, see: (a) Saitoh, A.; Achiwa, K.; Tanaka, K.; Morimoto, T. J. Org. Chem. 2000, 65, 4227. For two very recent reviews on amino acid-based phosphine catalysts, see: (b) Wang, T.; Han, X.; Zhong, F.; Yao, W.; Lu, Y. Acc. Chem. Res. 2016, 49, 1369. (c) Li, W.; Zhang, J. Chem. Soc. Rev. 2016, 45, 1657 and references therein. (23) CCDC 1546775 (2e), 1546776 (5a), and 1546777 (6e) contain the supplementary crystallographic data (Supporting Information). (24) The relative configurations of 6a and 6b were determined by NOE.

Program (C.-A.F.). We thank Prof. Jincai Wu (Lanzhou University) for X-ray crystallographic refinement of 6e.



REFERENCES

(1) (a) Sun, X.; Sneden, A. T. Planta Med. 1999, 65, 671. (b) Zhang, X.f.; Wang, H.-m.; Song, Y.-l.; Nie, L.-h.; Wang, L.-f.; Liu, B.; Shen, P.; Liu, Y. Bioorg. Med. Chem. Lett. 2006, 16, 949. (c) Hansen, U.; Schaus, S.; Grant, T.; Bishop, J.; Kavouris, J.; Christadore, L. M. PCT Int. Appl. WO2012050985A1, 2012. (d) Singh, P.; Faridi, U.; Srivastava, S.; Kumar, J. K.; Darokar, M. P.; Luqman, S.; Shanker, K.; Chanotiya, C. S.; Gupta, A.; Gupta, M. M.; Negi, A. S. Chem. Pharm. Bull. 2010, 58, 242. (e) Xu, S.; Shang, M.-Y.; Liu, G.-X.; Xu, F.; Wang, X.; Shou, C.-C.; Cai, S.-Q. Molecules 2013, 18, 5265. (2) (a) Ozturk, T.; McKillop, A. Can. J. Chem. 2000, 78, 1158. (b) Huh, J. R.; Englund, E. E.; Wang, H.; Huang, R.; Huang, P.; Rastinejad, F.; Inglese, J.; Austin, C. P.; Johnson, R. L.; Huang, W.; Littman, D. R. ACS Med. Chem. Lett. 2013, 4, 79. (3) For selected examples from ten recent years, see: (a) Sun, J.; Ding, W.-X.; Hong, X.-P.; Zhang, K.-Y.; Zou, Y. Chem. Nat. Compd. 2012, 48, 16. (b) Niharika, P.; Ramulu, B. V.; Satyanarayana, G. Org. Biomol. Chem. 2014, 12, 4347. (4) For review, see: Mieriņa, I.; Jure, M.; Stikute, A. Chem. Heterocycl. Compd. 2016, 52, 509. (5) (a) Chen, G.; Tokunaga, N.; Hayashi, T. Org. Lett. 2005, 7, 2285. (b) Korenaga, T.; Maenishi, R.; Osaki, K.; Sakai, T. Heterocycles 2010, 80, 157. (c) Luo, Y.; Carnell, A. J. Angew. Chem., Int. Ed. 2010, 49, 2750. (d) Mino, T.; Miura, K.; Taguchi, H.; Watanabe, K.; Sakamoto, M. Tetrahedron: Asymmetry 2015, 26, 1065. (6) (a) McGuire, M. A.; Shilcrat, S. C.; Sorenson, E. Tetrahedron Lett. 1999, 40, 3293. (b) Kim, H.; Yun, J. Adv. Synth. Catal. 2010, 352, 1881. (7) Li, X.-H.; Fang, P.; Chen, D.; Hou, X.-L. Org. Chem. Front. 2014, 1, 969. (8) For very recent selected examples, see: (a) Hu, H.; Liu, Y.; Guo, J.; Lin, L.; Xu, Y.; Liu, X.; Feng, X. Chem. Commun. 2015, 51, 3835. (b) Caruana, L.; Mondatori, M.; Corti, V.; Morales, S.; Mazzanti, A.; Fochi, M.; Bernardi, L. Chem. - Eur. J. 2015, 21, 6037. (c) Yu, X.-Y.; Chen, J.-R.; Wei, Q.; Cheng, H.-G.; Liu, Z.-C.; Xiao, W.-J. Chem. - Eur. J. 2016, 22, 6774. (d) Wang, Y.; Pan, J.; Dong, J.; Yu, C.; Li, T.; Wang, X.S.; Shen, S.; Yao, C. J. Org. Chem. 2017, 82, 1790. (e) Wu, B.; Yu, Z.; Gao, X.; Lan, Y.; Zhou, Y.-G. Angew. Chem., Int. Ed. 2017, 56, 4006. (f) Zhou, J.; Wang, M.-L.; Gao, X.; Jiang, G.-F.; Zhou, Y.-G. Chem. Commun. 2017, 53, 3531. (9) (a) Mahatthananchai, J.; Kaeobamrung, J.; Bode, J. W. ACS Catal. 2012, 2, 494. (b) Li, G.-T.; Gu, Q.; You, S.-L. Chem. Sci. 2015, 6, 4273. (c) Li, G.-T.; Li, Z.-K.; Gu, Q.; You, S.-L. Org. Lett. 2017, 19, 1318. (10) (a) Zhang, L.; Qureshi, Z.; Sonaglia, L.; Lautens, M. Angew. Chem., Int. Ed. 2014, 53, 13850. (b) Dong, C.; Alper, H. Tetrahedron: Asymmetry 2004, 15, 35. (11) Baeyer, A.; Villiger, V. Ber. Dtsch. Chem. Ges. 1903, 36, 2774. (12) For recent reviews on the chemistry of p-QMs, see: (a) Parra, A.; Tortosa, M. ChemCatChem 2015, 7, 1524. (b) Caruana, L.; Fochi, M.; Bernardi, L. Molecules 2015, 20, 11733. (c) Chauhan, P.; Kaya, U.; Enders, D. Adv. Synth. Catal. 2017, 359, 888. (13) For very recent selected examples on nonasymmetric 1,6-addition of p-QMs, see: (a) Zhang, Z.-P.; Dong, N.; Li, X. Chem. Commun. 2017, 53, 1301. (b) Jadhav, A. S.; Anand, R. V. Org. Biomol. Chem. 2017, 15, 56. (c) Ke, M.; Song, Q. Adv. Synth. Catal. 2017, 359, 384. (d) Molleti, N.; Kang, J. Y. Org. Lett. 2017, 19, 958. (e) Zhuge, R.; Wu, L.; Quan, M.; Butt, N.; Yang, G.; Zhang, W. Adv. Synth. Catal. 2017, 359, 1028. (f) Goswami, P.; Singh, G.; Anand, R. V. Org. Lett. 2017, 19, 1982. (g) Lin, C.; Shen, Y.; Huang, B.; Liu, Y.; Cui, S. J. Org. Chem. 2017, 82, 3950. (14) For catalytic asymmetric 1,6-addition of p-QMs, see: (a) Chu, W.D.; Zhang, L.-F.; Bao, X.; Zhao, X.-H.; Zeng, C.; Du, J.-Y.; Zhang, G.-B.; Wang, F.-X.; Ma, X.-Y.; Fan, C.-A. Angew. Chem., Int. Ed. 2013, 52, 9229. (b) Caruana, L.; Kniep, F.; Johansen, T. K.; Poulsen, P. H.; Jørgensen, K. A. J. Am. Chem. Soc. 2014, 136, 15929. (c) Lou, Y.; Cao, P.; Jia, T.; Zhang, Y.; Wang, M.; Liao, J. Angew. Chem., Int. Ed. 2015, 54, 12134. (d) Wang, Z.; Wong, Y. F.; Sun, J. Angew. Chem., Int. Ed. 2015, 54, 3210

DOI: 10.1021/acs.orglett.7b01331 Org. Lett. 2017, 19, 3207−3210